Download High-speed imaging of short wind waves by shape from refraction

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Gravitational lens wikipedia , lookup

Transcript
J. Europ. Opt. Soc. Rap. Public. 9, 14015 (2014)
www.jeos.org
High-speed imaging of short wind waves by shape from
refraction
D. Kiefhaber
S. Reith
Universität Heidelberg, Institut für Umweltphysik, INF 229, 69120 Heidelberg, Germany
Universität Heidelberg, Heidelberg Collaboratory for Image Processing (HCI)
Universität Heidelberg, Institut für Umweltphysik, INF 229, 69120 Heidelberg, Germany
Universität Heidelberg, Heidelberg Collaboratory for Image Processing (HCI)
R. Rocholz
Universität Heidelberg, Institut für Umweltphysik, INF 229, 69120 Heidelberg, Germany
now at: Heidelberg Engineering, Tiergartenstraße 15, 69121 Heidelberg, Germany
B. Jähne
Universität Heidelberg, Institut für Umweltphysik, INF 229, 69120 Heidelberg, Germany
Universität Heidelberg, Heidelberg Collaboratory for Image Processing (HCI)
[email protected]
This paper introduces the first high-speed system for slope imaging of wind-induced short water waves. The imaging slope gauge method is
used, which is based on the shape from refraction principle. The downward looking camera with a telecentric lens observes the refraction
of light rays coming from a high power custom telecentric LED light source that is placed underneath the wind wave facility. The light
source can be programmed to arbitrary intensity gradients in the x- and y-direction, so that the origin of a light ray is coded in intensity.
Four gradient images (acquired at 6000 fps) are combined for one 2D slope image. By only using intensity ratios, the measurements
become independent of lens effects from the curved water surface and inhomogeneities in the light source. Independence of wave height
is guaranteed by using telecentric illumination and telecentric imaging. The system is capable to measure the slopes of a wind-driven
water surface in the Heidelberg Aeolotron wind-wave facility on a footprint of 200 × 160 mm with a spatial resolution of 0.22 mm and a
temporal resolution of more than 1500 fps. For the first time, it is now possible to investigate the structure of short wind-induced waves
with sufficient spatial and temporal resolution to study their dynamic characteristics without aliasing effects. Example images and a video
of a 3D reconstructed water surface are shown to illustrate the principle.
[DOI: http://dx.doi.org/10.2971/jeos.2014.14015]
Keywords: Imaging slope gauge, shape from refraction, wave slope imaging, gravity-capillary waves, air sea interaction
1 INTRODUCTION
1.1 Short ocean waves
Short wind waves with wavelengths of less than a couple of
metres down to a few millimetres are important in many contexts. Surface roughness on this scale determines the strength
of radar backscatter [1], the distribution of the slopes of smallscale waves is important for (ocean) radiative transfer modelling [2, 3], they affect remote ocean colour measurements
[4]. They also play a prominent role as a major source of
near-surface turbulence, influencing the exchange of heat,
mass, and momentum between the atmosphere and the ocean
[5]–[7].
The measurement of these small-scale waves is challenging:
Their amplitudes are orders of magnitude smaller than those
of the longer waves, so that all amplitude-based measurement
instruments need a very high dynamic range. Surface slope on
the other hand has a limited range, since steep waves become
unstable and break. In fact, the slopes of very short (capillary)
waves, for which the water’s surface tension contributes significantly as a restoring force, are even higher than those of
longer waves.
A second effect of the surface tension has significant consequences for small-scale wave measurement: the water surface
is always smooth at the scale of optical wavelengths. Due to
this, all reflections are fully specular. This prohibits the use
of standard surface reconstruction techniques (shape from
stereo, shape from shading) that have been developed for
Lambertian surfaces [8]. Specialized measurement techniques
have been developed that use the dependence of the reflection
coefficient on the incidence angle of light [9] (known as Stilwell photography in oceanography or deflectometry in industrial
inspection [10, 11]). Unfortunately, the reflectance of the
water surface is highly non-linear. For incidence angles of
0°-40°, it barely changes, above Brewster’s angle (53°) it
increases rapidly. In addition, very homogeneous (or well
known) illumination is required. An extension of the technique that uses the degree and orientation of polarization
of reflected light for slope measurements has recently been
developed [12]. The dependence of these quantities on the
incidence angle is more linear than that of the reflection coefficient. The illumination is not required to be homogeneous,
since only the ratio of measured intensities is evaluated, but it
needs to be fully unpolarized (or have known polarization).
Received November 05, 2013; revised ms. received January 27, 2014; published March 24, 2014
ISSN 1990-2573
J. Europ. Opt. Soc. Rap. Public. 9, 14015 (2014)
D. Kiefhaber, et al.
For laboratory applications, where mounting parts of the instrument under water is possible, refraction-based measurements are superior to reflection-based techniques for a number of reasons: Only 2% of the light falling on the water surface are reflected, so that very high luminance light sources are
required for reflection measurements. Because the angles are
doubled at the point of reflection, these measurements also require a much larger light source. Also, the dependence of total
ray deflection on surface slope is less non-linear for refraction
measurements [8].
mirror
f = 2.0 m
telecentric lens
window
1.4 m
1.0 m
window
1.2 Shape from Refraction in Wave Imaging
A slope meter based on the refraction of light at the water surface was first used by Cox [13]. With a telescope and a photometer he observed light refracted at the water surface. The
origin of the light rays could be determined because the light
source intensity had a spatial gradient. Unfortunately, in this
simple setup the measured intensity was also depending on
wave height and surface curvature.
Fresnel lens
f = 0.76 m
light source
FIG. 1 Wave imaging setup at the Heidelberg circular (annular) wind-wave facility.
air
α
90°
The measurement scheme was inverted in the 1970s, when
the laser slope gauge was developed [14]–[16]: The refraction
of a vertical laser beam falling down onto the water surface is
tracked with an underwater camera observing a screen. Measurements are independent of curvature and, if the screen is
placed in the focal plane of a lens, of water height. Still, measurements were limited to single points on the water surface,
although attempts were made to scan the laser over an area
on the water surface [17].
Real 2D measurements required an extension of Cox’s technique to the imaging slope gauge: A camera is looking down
onto the water surface, into refracted light coming from an underwater light source with one or two spatial gradients. This
way, it is possible to measure surface slope on a two dimensional footprint. Using a camera and an area light source with
one or two spatial gradients, it is possible to measure both
slope components on a two-dimensional area [18]–[22]. All
previous devices could, however, acquire wave slope image
sequences at most at video rates with 60 fps. The purpose of
this paper is to describe the first optical setup for a wave slope
imaging instrument with a) telecentric imaging and illumination and b) a computer-controlled four-gradient LED light
source, which is bright enough for high-speed imaging with
up to 1500 fps, so that even capillary waves with the highest
frequencies can be sampled without aliasing.
α
β
water
water surface
γ
γ
bottom window
air
δ
Fresnel lens
δ
f
air
illumination
screen
principal axis
of Fresnel lens
FIG. 2 Ray geometry of the imaging slope gauge.
Both the camera and the water surface are placed in the two
focal planes of a large custom BK7 planoconvex lens (diameter
d = 0.32 m, focal length f = 2 m). Together with a standard
105 mm camera lens, the two lenses form an object-space telecentric lens, ensuring that all light rays that enter the camera
are parallel on the right side of the lens. They are also orthogonal to the mean water surface. As a consequence, the magnification factor is independent of the distance. This is of importance, since wave heights in the Aeolotron facility can exceed
60 cm. At a distance of 4 m, this would lead to changes in the
size of structures on the water surface of up to 25%.
The fact that all light rays are orthogonal to the mean water
surface is also beneficial for the illumination design. In the
absence of diverging rays above the water surface, the light
source can be more compact without a loss of measurement
range.
2 THE HIGH-SPEED IMAGING SLOPE
GAUGE
2.1 Imaging setup
Figure 1 is a schematic of the setup that is used at the Heidelberg Aeolotron wind-wave facility. A high-speed camera
is observing the water surface from a distance of 4 m. The water body is illuminated from below with a programmable high
power LED light source. On its way to the camera, the light is
refracted at the surface.
Below the water body, an 89 cm diameter Fresnel lens
(f = 76.2 cm) is used to guarantee telecentric illumination. To
understand this, it is helpful to trace a ray from the camera to
the light source, as shown in Figure 2: A ray of light penetrating the water vertically from above is refracted at the water
surface. For an incidence angle α to the surface normal, the
14015- 2
J. Europ. Opt. Soc. Rap. Public. 9, 14015 (2014)
D. Kiefhaber, et al.
in the x and y direction are computed from two opposing intensity gradients each. This ratio imaging corrects for lens effects at the water surface, but also significantly reduces the
demands to the homogeneity of the light source.
36 LEDs per row
A
B
C
1 cm
24 LEDs per column
D
FIG. 3 Sketch of light source; for the x direction, there are 35 columns with 24 LEDs
each, for the y direction there are 24 rows with 36 LEDs each. Custom electronics
control the LED current of each row/column. Four different brightness gradients (A-D)
are pre-programmed.
refracted angle is given by Snell’s law na sin(α) = nw sin( β),
where na ≈ 1 and nw ≈ 1.33 are the refractive indexes of air
and water, respectively. Further refractions occur entering
and leaving the polycarbonate bottom window (n = 1.58) of
the wind-wave facility.
Because the light source is placed in the focal plane of the Fresnel lens, all rays leaving the water at an angle δ to the vertical
are focused onto the same position of the illumination screen.
This effectively translates different water surface slopes into
unique screen positions, independent of position on the water
surface.
2.2 Position coding
A light source for the measurement of only one slope component can simply have a gradient in the direction in which
slope is to be measured. An extension for two slope components is possible by using a colour-coded light source and
an RGB camera [23]. In this scheme, the third colour channel
can then be used for normalization and the correction of curvature (lens) effects at the water surface, e.g. by demanding
that the total intensity I = IR + IG + IB is constant [24, sec.
8.5.3]. Unfortunately, this leads to significant problems with
chromatic aberration of the large telecentric lens. In our new
setup, a different approach is chosen: The light source consists of columns and rows of LEDs wired in series (see Figure 3). Each row/column can be programmed to two different brightness settings using custom electronics that can control current from 0 - 1000 mA with 12 bit resolution and fast
multiplexers to swap the two states. In total, the light source
consists of 1704 red high power LEDs (Cree XLamp XP-E
red, peak wavelength 630 nm, max. cont. current 700 mA). The
LED spacing is 2.1 cm in the rows/columns. An acrylic diffusing screen placed at a distance of 4 cm is used to homogenize
the luminance gradient.
With this light source, and the four brightness gradients A - D
shown in Figure 3, the following intensity coding scheme can
be implemented: Normalized intensities
Ix =
A−B
A+B
and
Iy =
C−D
,
C+D
(1)
To test the performance and measurement accuracy of the system, a float with a planoconvex BK7 glass lens was put on the
water surface (Figure 4(a)). From the known curvature radius
of the lens, it is possible to compute the surface slope tan α0 for
every position on the lens. An “equivalent water slope” can be
computed from Snell’s law (see equations in Figure 4(a)) by
setting γ = γ0 . For small angles, the equivalent water slope is
1.58 times the slope on the BK7 lens.
In Figure 4(b), a raw image of the lens float (illumination gradient C) is shown; the axis labels are in units of slope (tan α0 ).
Even though a diffusing screen is used to homogenize the
light source, the individual LEDs can still be distinguished.
Also, a number of dents in the bottom window of the facility are visible as darker spots, e.g. at position s x = 0.1,
sy = −0.25. In Figure 4(c), the corresponding normalized image (see Eq. 1) is shown. The inhomogeneities have disappeared; a smooth and monotonous gradient corresponding to
the slope of the lens in the y direction is visible.
Figure 4(d) shows a profile of the 2D distribution in Figure 4(c). The slope in the y-direction is shown as a function
of the normalized intensity (which was coded in grey values
in Figure 4(c)). The red curve shows the slope of the lens on
the calibration float (tan α0 ), the blue curve the corresponding water slope that was computed using the considerations
in Figure 4(a) (tan α). The shaded areas indicate the standard
deviation of individual normalized images around the mean.
The average error of the slope measurement is 0.0251 .
2.3 High-speed, high-resolution camera
To avoid aliasing when imaging fast-moving small-scale
waves, good resolution is required in both the time and
space domain. In our setup, a pco.dimax camera (PCO AG)
is used. It allows acquisition of up to 6030 fps at a spatial
resolution of 960 × 768 pixels. Since four raw images are
required to compute a 2D wave slope image, the effective
frame rate for slope measurements is 1507.5 fps. The footprint
on the water surface is 203 × 166 mm, corresponding to a
pixel size of 0.22 mm. For each image, the integration time is
140 µs and the LEDs are flashed accordingly. The brightness
is sufficient for measurements with f-number 8 (camera lens
with f = 105 mm).
3 RESULTS
The imaging slope gauge was set up at the Heidelberg Aeolotron wind-wave facility in May 2013. Some processed measurements demonstrating the capabilities of the instrument
are shown in the video clip added as supplementary material
and represented by Figure 5.
1 Very recently, the maximum brightness of the light source has been
tripled. In future measurements, the error will be smaller.
14015- 3
J. Europ. Opt. Soc. Rap. Public. 9, 14015 (2014)
D. Kiefhaber, et al.
a
b
air
α‘
β‘
−0.5
BK7 lens
d
δ‘
R
tan β‘ = tan α‘ / nG
δ‘ = α‘ - β‘
tan γ‘ = tan δ‘ nW /nG
α‘
γ‘
water
−0.25
0
air
α
0.25
β
γ
0.5
tan β = tan α / nW
γ=α-β
−0.5
water
−0.25
0
0.25
0.5
d
c
1
BK7 lens
water
−0.5
0.5
slope
−0.25
0
0.25
0
−0.5
0.5
−0.5
−0.25
0
0.25
−1
−0.5
0.5
0
normalized intensity
0.5
FIG. 4 (a) A float with a BK7 lens is used for validation. The slope tan α0 can be converted into the equivalent water slope tan α by applying Snell’s law to both scenarios and
setting γ0 = γ. (b) Raw image of a lens floating on the water surface, axes are labelled in units of slope tan α0 . (c) Normalized intensity Iy (see Eq. (1)). (d) The dependence of
slope on normalized intensity (red); profile along the red line in c. Also shown is the corresponding equivalent water slope (blue).
The video shows a rendering of the reconstructed 3D surface
with software developed by [25]. The necessary conversion
from slope to height is done using an algorithm proposed by
[26]. The Fourier transform of the height
ĥ(k x , k y ) = −i
k x ŝ x + k y ŝy
k2x + k2y
!
,
(2)
with the spatial frequency (wavenumber) k x , k y , and the
Fourier transforms of the slope images ŝ x , ŝy . After applying the inverse Fourier transform to the right hand side of
Eq. (2), a surface elevation map is obtained. Due to the integration, changes in the mean elevation, e.g. by waves with
wavelengths larger than the footprint of the image, are lost.
This drawback of wave slope imaging can be made up for by
additionally measuring wave height at a single point in the
image footprint.
The video shows a typical wave field at medium wind speeds,
shown 50 times slower than real time. Characteristic short
gravity waves (with wavelengths of about 10 cm) are visible,
on their leeward side, so-called parasitic capillary waves are
formed. These small ripples are created from an instability of
the short gravity wave and propagate as bound waves with
the same phase velocity as their carrier wave.
In Figure 6, a breaking wave is entering the footprint from the
right side. This leads to problems in the reconstruction, since
a) multiple refractions can occur when air is entrained in the
water or spray is created in the air, and b) surface slopes can
become very large, so that the refracted rays miss the light
source. In our setup, slopes up to 26.3° (tan α = 0.5) in the
alongwind and 19.6° (0.37) in the crosswind direction can be
measured.
While these 3D visualizations are a useful tool to observe
the dynamics of the wavy water surface, quantitative physical information is typically analysed with the help of 3D
wavenumber-frequency power spectra, which contain the distribution of energy to the different wavelengths, as well as the
dispersion relation ω (k ) . Wavelengths of individual waves
can also be extracted with spatial band-pass filtering (e.g. Gabor filters, see [24], sec. 13.4.5).
The complex spatio-temporal evolution of a wind wave field
can also be studied with the help of space-time images (Figure 7). Each column of the image contains a slope profile
of the water surface, along a line (of length 20 cm) in the
wind direction. The slope is coded in grey values, white
is a high slope away from the wind, black is towards the
14015- 4
J. Europ. Opt. Soc. Rap. Public. 9, 14015 (2014)
D. Kiefhaber, et al.
FIG. 7 Space-time image of the wave field at a wind speed of 4.2 m/s. Each column
contains a snapshot of the slope along a 20 cm transect of the surface. The image
shows the temporal evolution of the slope; time increases to the right (330 ms total),
the wind is blowing from top to bottom. Slope is coded in grey values, with slopes
away from the wind shown in white.
FIG. 5 3D rendering of the wave field in the Aeolotron facility at 6.4 m/s. Characteristic
short gravity waves with parasitic capillaries (the small ripples) are visible. Video
ing to yield the (modulated) velocity of the waves (e.g. [24],
sec. 14.2.4).
4 CONCLUSION AND OUTLOOK
We have installed the first high speed system for measuring the dynamics of small-scale water waves. The instrument
captures the two-dimensional distribution of surface slope
with a rate of more than 1500 fps and a spatial resolution of
0.22 mm. With this system, measurements of 3D frequencywavenumber spectra of short waves down to the smallest ripples are possible for the first time without aliasing. With this
system, we can study the dynamics and interactions of smallscale waves, e.g. deviations from the linear dispersion relation
or non-linear interactions.
FIG. 6 At higher wind speeds (8.5 m/s), breaking waves cause problems: Multiple
refraction of the light occurs at bubbles and spray, large slopes outside the measure-
At the same footprint on the water surface, we can also visualize the horizontal structure of the mass boundary layer, allowing us to investigate interactions between waves, near-surface
turbulence and gas exchange [27].
ment range are common.
5 ACKNOWLEDGEMENTS
wind. In the space time image, the wind direction is from top
to bottom. The successive columns of the space-time image
show the time evolution of the slope profile over a span of
330 ms.
A (microscale) breaking short gravity wave is passing through
the ISG footprint (yellow/1). Parasitic capillary waves (very
small-scale ripples) that are created on the leeward side are
clearly visible. The breaking process is more subtle than with
large scale waves; no bubbles are created in the process. After
the microbreaker has passed, the water surface is flat, waves
are destroyed by the turbulence created in the breaking process. Not all waves travel in the direction of the wind; parallel
to the red arrow (2) weak structures are visible that propagate
against the wind direction. The modulation of the propagation
of smaller waves by the orbital velocity of the dominant larger
waves is also visible (blue/3). The blue arrows, which follow
the local space-time gradient clearly have different slopes. The
local space-time gradient can be extracted by image process-
This work is based on a presentation at the “1st EOS Topical Meeting on Frontiers of Optical Imaging” held in Murten
(CH) in September 2013. Partial financial support for the
project by the German Federal Ministry of Education and Research (BMBF) joint project “Surface Ocean Processes in the
Anthropocene” (SOPRAN, FKZ 03F0611F) is gratefully acknowledged.
References
[1] G. R. Valenzuela, “Theories for the interaction of electromagnetic
and ocean waves - A review,” Bound.-Lay. Meteorol. 13, 61–85
(1978).
[2] Z. Jin, T. P. Charlock, K. Rutledge, K. Stamnes, and Y. Wang, “Analytical solution of radiative transfer in the coupled atmosphereocean system with a rough surface,” Appl. Optics 45, 7443–7455
(2006).
14015- 5
J. Europ. Opt. Soc. Rap. Public. 9, 14015 (2014)
D. Kiefhaber, et al.
[3] M. Hieronymi, “Monte Carlo code for the study of the dynamic
light field at the wavy atmosphere-ocean interface,” J. Europ. Opt.
Soc. Rap. Public. 8, 13039 (2013).
[17] E. J. Bock and T. Hara, “Optical Measurements of Ripples Using
a Scanning Laser Slope Gauge,” Proc. SPIE - Optics of the Air-Sea
Interface 1749, 272 (1992).
[4] H. R. Gordon and M. Wang, “Retrieval of water-leaving radiance
and aerosol optical thickness over the oceans with SeaWiFS: a
preliminary algorithm,” Appl. Optics 33, 443–452 (1994).
[18] W. C. Keller and B. L. Gotwols, “Two-dimensional Optical Measurement of Wave Slope,” Appl. Optics 22, 3476–3491 (1983).
[5] B. Jähne, K. O. Münnich, R. Bösinger, A. Dutzi, W. Huber, and P. Libner, “On the parameters influencing air-water gas exchange,” J.
Geophys. Res. 92, 1937–1950 (1987).
[6] N. M. Frew, E. J. Bock, U. Schimpf, T. Hara, H. Haußecker, J. B. Edson, W. R. McGillis, et al., “Air-sea gas transfer: Its dependence
on wind stress, small-scale roughness, and surface films,” J. Geophys. Res. 109, C08S17 (2004).
[7] R. Wanninkhof, W. E. Asher, D. T. Ho, C. Sweeney, and
W. R. McGillis, “Advances in quantifying air-sea gas exchange and
environmental forcing,” Annu. Rev. Mar. Sci. 1, 213–244 (2009).
[8] B. Jähne, J. Klinke, and S. Waas, “Imaging of short ocean wind
waves: a critical theoretical review,” J. Opt. Soc. Am. 11, 2197–2209
(1994).
[19] B. Jähne, Transfer processes across the free water interface (Habilitation thesis, Institut für Umweltphysik, Fakultät für Physik und
Astronomie, University Heidelberg, 1985).
[20] B. Jähne and K. Riemer, “Two-dimensional wave number spectra of small-scale water surface waves,” J. Geophys. Res. 95,
11531–11646 (1990).
[21] X. Zhang and C. S. Cox, “Measuring the two-dimensional structure
of a wavy water surface optically: A surface gradient detector,”
Exp. Fluids 17, 225–237 (1994).
[22] G. Balschbach, J. Klinke, and B. Jähne, “Multichannel shape from
shading techniques for moving specular surfaces,” in Proceedings
of 5th European Conference on Computer Vision (ECCV), 170–184
(Freiburg, 1998).
[9] D. J. Stilwell, “Directional energy spectra of the sea from photographs,” J. Geophys. Res. 74, 1974–1986 (1969).
[23] R. Rocholz, Spatiotemporal Measurement of Short Wind-Driven Water Waves (Dissertation, Institut für Umweltphysik, Fakultät für
Physik und Astronomie, University Heidelberg, 2008).
[10] O. Kafri and I. Glatt, “Moire Deflectometry: A Ray Deflection Approach To Optical Testing,” Opt. Eng. 24, 246944–246944– (1985).
[24] B. Jähne, Digital Image Processing (Sixth Edition, Springer, Berlin,
2005).
[11] M. Knauer, J. Kamisnski, and G. Häusler, “Phase Measuring Deflectometry: a new approach to measure specular free-form surfaces,”
Proc. SPIE - Optical Metrology in Production Engineering 5457,
366–376 (2004).
[25] S. Wanner, C. Sommer, R. Rocholz, M. Jung, F. A. Hamprecht, and
B. Jähne, “A Framework for Interactive Visualization and Classification of Dynamical Processes at the Water Surface,” in 16th
International Workshop on Vision, Modelling and Visualization,
K. P. Peter Eisert, Joachim Hornegger, eds., 199–206 (Eurographics Association, Goslar, 2011).
[12] C. Zappa, M. Banner, H. Schultz, A. Corrada-Emmanuel, L. Wolff,
and J. Yalcin, “Retrieval of short ocean wave slope using polarimetric imaging,” Meas. Sci. Technol. 19, 055503 (13pp) (2008).
[13] C. S. Cox, “Measurement of slopes of high-frequency wind waves,”
J. Mar. Res. 16, 199–225 (1958).
[14] R. Tober, R. Anderson, and O. Shemdin, “Laser instrument for detecting water ripple slopes,” Appl. Optics 12, 788–794 (1973).
[15] J. C. Scott, “An optical probe for measuring water wave slopes,” J.
Phys. E Sci. Instrum. 7, 747–749 (1974).
[26] R. T. Frankot and R. Chellappa, “A Method for Enforcing Integrability in Shape from Shading Algorithms,” IEEE T. Pattern Anal. 10,
439–451 (1988).
[27] C. Kräuter, D. Trofimova, D. Kiefhaber, N. Krah, and B. Jähne, “HighResolution 2-D Fluorescence Imaging of the Mass Boundary Layer
at Free Water Surfaces,” to be published in J. Europ. Opt. Soc. Rap.
Public. (2014).
[16] B. A. Hughes, R. W. Grant, and R. W. Chappell, “A fast response
surface-wave slope meter and measured wind-wave moments,”
Deep-Sea Res. 24, 1211–1223 (1977).
14015- 6